You are on page 1of 9

Research Article

www.acsami.org

Mg(OH)2 Supported Nanoscale Zero Valent Iron Enhancing the


Removal of Pb(II) from Aqueous Solution
Minghui Liu,†,∥ Yonghao Wang,†,‡,∥ Luntai Chen,†,∥ Yan Zhang,†,∥ and Zhang Lin*,†,§,∥

Key Laboratory of Design and Assembly of Functional Nanostructures, Fujian Institute of Research on the Structure of Matter,
Chinese Academy of Sciences, Fuzhou, Fujian 350002, China

College of Environment and Resources, Fuzhou University, Fuzhou, Fujian 350002, China
§
School of Environment and Energy, South China University of Technology, Guangzhou 510006, China

Fujian Provincial Key Laboratory of Nanomaterials, Fujian Institute of Research on the Structure of Matter, Chinese Academy of
Sciences, Fuzhou, Fujian 350002, China
*
S Supporting Information

ABSTRACT: In this article, a novel composite (Mg(OH)2


supported nanoscale zerovalent iron (denoted as nZVI@Mg-
(OH)2) was prepared and characterized by X-ray diffraction,
scanning electron microscopy, and transmission electron micros-
copy method. The morphology analysis revealed that Mg(OH)2
appeared as self-supported flower-like spheres, and nano Fe0
particles were uniformly immobilized on the surface of their
“flower petals”, thus aggregation of Fe0 particles was minimized.
Then the Pb(II) removal performance was tested by batch
experiments. The composite presented exceptional removal
capacity (1986.6 mg/g) compared with Mg(OH)2 and nanoscale
zerovalent iron due to the synergistic effect. Mechanisms were also
explored by a comparative study of the phase, morphology, and
surface valence state of composite before and after reaction, indicating that at least three paths are involved in the synergistic
removal process: (1) Pb(II) adsorption by Mg(OH)2 (companied with ion exchange reaction); (2) Pb(II) reduction to Pb0 by
nanoscale zerovalent iron; and (3) Pb(II) precipitation as Pb(OH)2. The hydroxies provided by Mg(OH)2 can dramatically
promote the role of nanoscale zerovalent iron as reducer, thus greatly enhancing the whole Pb(II) sequestration process. The
excellent performance shown in our research potentially provides an alternative technique for Pb(II) pollution treatment.
KEYWORDS: nanoscale zerovalent iron (nZVI), magnesium hydroxide, nZVI@Mg(OH)2 composite, synergistic sequestration

■ INTRODUCTION
Lead can cause acute, chronic circulatory, neurological,
removing various pollutants, such as chlorinated organic
compounds,9 nitroaromatic compounds,10 dyes,11 and heavy
hematological, gastrointestinal, reproductive, and immunolog- metal ions,12−15 including Pb(II).16−21 Nevertheless, nZVI
ical pathologies,1−3 because of its toxicity and nonbiodegrad- particles tend to aggregate due to the high surface energy and
ability. In China, some industrial processes including lead magnetic force, leading to sharp deterioration of activity,
mining, smelting, and consuming can discharge lead-bearing durability, and efficiency, which would greatly hinder the large-
wastewater, especially in the lead-acid batteries industry, where scale application of nZVI.22 Incorporating nZVI particles onto
over 20 million tons of wastewater are generated per year, support materials can successfully ease the above-mentioned
containing more than 150 tons of Pb(II). Therefore, a cost- problems. The commonly used support materials include
effective method must be developed to treat Pb(II)-bearing kaolinite,11,17 zeolite,20 clay,23,24 bentonite25 and chelating
wastewater. resin,16 mesoporous carbon,10,14 silica,9 and graphene,26
At present, extensive methods have been employed to which all can reduce the aggregation of nZVI particles, thus
remedy heavy metal polluted wastewater, including chemical improving their performance in pollutant treatment. However,
precipitation, ion-exchange, membrane separation, and adsorp- some shortcomings of the reported support materials, such as
tion.4−6 Among them, adsorption is considered as one of the complexity of preparation, high cost, and uneven loading, still
most attractive ways due to its easy processability and high limit the wide application of nZVI. Hence, the ideal carrier to
efficiency. In the past two decades, nanoscale zerovalent iron
(nZVI) has received much attention because of its large specific Received: December 30, 2014
surface area, high reaction activity, and strong reductive Accepted: March 31, 2015
power.7,8 Studies show that nZVI appears to be effective in Published: March 31, 2015

© 2015 American Chemical Society 7961 DOI: 10.1021/am509184e


ACS Appl. Mater. Interfaces 2015, 7, 7961−7969
ACS Applied Materials & Interfaces Research Article

fulfill the potential of nZVI must be cost-effective, mechanically NaBH4 was added into the flask through titration with a rate of 3 mL/
stable, and compatible with nZVI particles. Especially, an min. Before and throughout the procedure, N2 was introduced to expel
intense affinity of support material toward target ions will the dissolved oxygen and keep an anaerobic environment. The flask
definitely enhance the performance of nZVI as a whole. was stirred at 300 r/min using a mechanical stirrer, and the stirring was
continued for another 30 min after finishing the titration. The resulted
In recent years, nanoscale magnesium hydroxide (Mg(OH)2) black paste was collected and washed with nanopure water and
has aroused extensive attention because of low-cost and absolute ethanol three times respectively, then dried in a vacuum oven
environmental friendliness. Researches indicate that Mg(OH)2 at 45 °C overnight, and stored for use. A composite with different Fe
exhibits a promising prospect in acid wastewater treatment, content was synthesized through varying the ratio of Mg(OH)2 and
heavy metal deprivation,27 and decoloration of printing and FeSO4·7H2O (the Fe mass fraction of composite used in following
dyeing wastewater. For instance, Wang et al.28 demonstrated experiments was 53.9% unless specially specified).
that nanoscale self-supported Mg(OH)2 aggregates could Unsupported nZVI was also prepared via the same procedure for
absorb low concentration anionic dye from water selectively. comparative study.
Li et al.29 reported that self-supported flowerlike Mg(OH)2 Characterizations and Measurements. X-ray diffraction
(XRD), scanning electron microscopy (SEM), transmission electron
could recycle rare earth ion from low concentration wastewater microscopy (TEM), and BET N2 adsorption method were employed
via ion-exchange. All those researches reveal that the to characterize the as-prepared Mg(OH)2, nZVI, and nZVI@
morphology of Mg(OH)2 can be easily controlled and modified Mg(OH)2 composite.32 XRD analysis was conducted on a PANalytical
to obtain a self-supported flowerlike structure with large specific X’Pert PRO diffractometer with Cu KR radiation in the continuous
surface area, which can potentially provide an ideal carrier for scanning mode at 40 kV, 40 mA from 2θ at 5 to 85° with a step size of
nZVI particles.30 Based on the potential attraction of Mg(OH)2 0.017°, and collection time of 20 s per step. SEM and TEM images
to Pb(II), we expected the combination of nZVI and Mg(OH)2 were obtained on a JSM-6700F scanning electron microscopy
could allow a synergy more than reducing aggregation of nZVI equipped with an Oxford-INCA energy dispersive X-ray (EDS)
particles, which would facilitate decontamination of Pb(II). spectroscopy and a JEOL JEM2010 transmission electron microscopy
at 200 kV. BET surface area of the materials was measured with an
In this article, we aim to develop a novel composite by ASAP 2020 specific surface area and porosity analyzer (Micromeritics
loading nZVI onto Mg(OH)2 (nZVI@Mg(OH)2). The Instrument Corp. U.S.).
performance of removing Pb(II) was investigated by batch Pb(II), Fe(II/III), and Mg(II) concentration in the liquid samples
experiments. In addition, the influence of Fe loading ratio, pH collected from batch experiment was determined by atomic absorption
value, and initial Pb(II) concentration on removal efficiency spectrometer (AAS, TAS-990FG, Purkinje General Analytical Instru-
were also tested. The results show that the agglomeration of ment corp. China), while Pb, Fe, and Mg element content of solid
nZVI particles was eliminated, and a superb capacity for Pb(II) samples as well as Fe mass fraction of raw composites was determined
uptake was achieved (1986.6 mg/g), much higher than the by inductively coupled plasma-atomic emission spectrometry (ICP-
weighted sum value (1246 mg/g) of its component, indicating AES, Jobin Yvon Ulima2).
Pb(II) Removal: Batch Experiments. The optimal experimental
the existence of a synergy between nZVI and Mg(OH)2. conditions were determined beforehand, details of the procedure were
Further mechanism analyses attested that Mg(OH)2 played a described in the Supporting Information. Batch experiments for Pb(II)
crucial role by releasing OH− which can enhance the reduction removal, including performances of different materials and effects of
of Pb(II). different initial Pb(II) concentration were conducted in a 250 mL

■ EXPERIMENTAL SECTION
Materials and Chemicals. All chemicals used in this study,
three-necked flask equipped with a mechanical stirrer at a speed of 300
r/min. The whole set of experiments were conducted at room
temperature and N2 environment (N2 was introduced 10 min before
the experiment started to expel the dissolved oxygen and throughout
including magnesium sulfate (MgSO4·7H2O), sodium borohydride
(NaBH4), lead acetate (Pb[CH3COO]2·3H2O), ferrous sulfate the procedure to keep an anaerobic environment).
(FeSO4·7H2O), sodium hydroxide, acetic acid (CH3COOH), and In order to compare the performance of Pb(II) removal by
nitric acid (HNO3), were analytic grade purchased from Sinopharm Mg(OH)2, nZVI, and nZVI@Mg(OH)2, 50 mg of each was added into
Chemical Reagent Shanghai Co. Ltd. and were used directly as 100 mL of continuously stirred Pb(CH3COO)2 stock solution of 1000
received without further purification. All solutions were prepared using mg/L. 3 mL of suspension was extracted at certain time intervals and
nanopure water (conductivity = 18 μΩ/m, Toc < 3 μg/L). A stock centrifuged to collect both liquid and solid samples. The liquid sample
solution of Pb(II) was prepared by dissolving 1.8306 g of lead acetate was used for Pb(II), Fe(II/III), and Mg(II) concentration analysis
(Pb[CH3COO]2·3H2O) into 1000 mL of nanopure water at a after being filtered by a syringe filter (0.22 μm) and acidized by HNO3
volumetric flask, and a Pb(II) solution of different concentrations was (0.01 M). The solid sample was dried for Pb, Fe, and Mg element
made by dilution of the stock solution. content analysis. The pH value of suspension was monitored and
Preparation of Mg(OH)2, nZVI and nZVI@Mg(OH)2 Compo- recorded every 1 min during the experiments.
site. Mg(OH)2 was prepared through facile precipitation method at The effect of the initial Pb(II) concentration was tested by the same
room temperature and atmospheric pressure. Briefly, NaOH (1 M) procedure except the initial concentration was altered to 250, 500, 750,
solution was introduced into the same volume of MgSO4 (0.5M) and 1000 mg/L.
C3H8O3/H2O hybrid solution (15:85, v/v) dropwise, while stirring To investigate the feasibility toward actual industrial Pb(II)-bearing
vigorously through the titration, and it was continuously stirred for wastewater, batch experiments were also conducted in acid Pb−Zn
another 6 h after the titration. After aging for 24 h, the precipitate was mine tailing leachate which was analogous to industrial drainage in
collected by centrifugation and followed by washing with nanopure terms of strong acidic and high multi-ion concentration solution
water and absolute ethanol several times and then dried at 50 °C for environment. The method was the same as the optimization
use. experiment.
A modified liquid phase sodium borohydride reduction method31
was employed to prepare Mg(OH)2 supported nZVI. In a typical
procedure, 0.416 g of Mg(OH)2 was dispersed in 75 mL of absolute
ethanol (ultrasonic dispersion for 5 min), 2.085 g of FeSO4·7H2O was
■ RESULTS AND DISCUSSION
Characterization of the Materials. Figure 1 demonstrates
dissolved in 75 mL of nanopure water, and then the suspension and XRD patterns of nZVI, Mg(OH)2, and nZVI@Mg(OH)2. As
Fe(II) solution were mixed in a 500 mL three-necked flask to make a shown in Figure 1, the peaks of nZVI are well indexed as ferrite
Fe(II) concentration of 0.05 M. Then the same volume of 0.25 M (JCPDS: 00-006-0696), of which 2θ at 44.67, 65.02, and 82.33°
7962 DOI: 10.1021/am509184e
ACS Appl. Mater. Interfaces 2015, 7, 7961−7969
ACS Applied Materials & Interfaces Research Article

equation. The XRD pattern of as-prepared Mg(OH)2 is well


indexed as brucite (JCPDS: 00-007-0239), with a calculated
crystallite size of 21.6 ± 1.1, 31.1 ± 1.5, and 53.9 ± 2.7 nm in
the [001], [101], and [110] directions, respectively.29 However,
for nZVI@Mg(OH)2, only weak peaks at 18.53, 37.98, and
58.67° can be recognized, which are coincident with [001],
[101], and [110] diffraction peaks of Mg(OH)2. Besides, a peak
at 44.67° seems to be formed by broadening the [110] peak of
iron, probably due to the low crystallinity degree of iron, which
is similar to the previous literature.31 In addition, the specific
surface area is 11.6, 41.3, and 40.2 m2/g for nZVI, Mg(OH)2,
and nZVI@Mg(OH)2, respectively. The high specific surface
Figure 1. XRD patterns of nZVI, nZVI@Mg(OH)2, and Mg(OH)2 area is almost maintained after 53.9% Fe loading (determined
(●, Fe; ■, Mg(OH)2) .
by ICP-AES), suggesting a well dispersion of nZVI particles.
Figure 2 shows the typical morphology of Mg(OH)2, nZVI,
corresponds to [110], [200], and [211] directions of α-Fe, and nZVI@Mg(OH)2. As shown in Figure 2a, as-prepared
respectively, indicating that unsupported nZVI has high purity Mg(OH)2 presents flower-like spheres with uniform size
and good crystallographic degree.32,33 The mean grain size of around 10 μm. Enlarged micrograph (Figure S1) reveals that
nZVI is calculated to be 22.6 ± 1.1 nm using the Scherrer the spheres are formed by interwoven Mg(OH)2 plates with a

Figure 2. (a) SEM image of flower-like Mg(OH)2 spheres; (b) SEM image of nZVI@Mg(OH)2; (c) TEM image of nZVI@Mg(OH)2; (d) TEM
image of nZVI; (e) HAADF-STEM image of nZVI@Mg(OH)2; and corresponding EDS elemental mapping images.

7963 DOI: 10.1021/am509184e


ACS Appl. Mater. Interfaces 2015, 7, 7961−7969
ACS Applied Materials & Interfaces Research Article

Figure 3. (a) Comparison of Pb(II) removal by nZVI@Mg(OH)2, nZVI, and Mg(OH)2 (dose: 500 mg/L, stirring speed: 300 r/min, initial
concentration: 1000 mg/L, initial pH: 6.86, temperature: 25 °C); (b) comparison of Pb(II) removal capacity and efficiency by nZVI@Mg(OH)2 and
weighted sum of the value by nZVI and Mg(OH)2 alone.

thickness of 10−20 nm. Morphology of nZVI@Mg(OH)2 in where QnZVI and QMg(OH)2 are the uptake amounts of Pb(II) by
Figure 2b,c reveals that nZVI particles are uniformly nZVI and Mg(OH)2 under the same experimental conditions.
immobilized on the surface of Mg(OH)2 plates, and nZVI wt %nZVI, wt % Mg(OH)2 are weight percentages of nZVI and
particles loaded on Mg(OH)2 have a diameter of 40−60 nm
Mg(OH)2 in the nZVI@Mg(OH)2 composite. As shown in
and a shell thickness of several nanometers.19,34 However,
Figure 3b, the values of capacity and efficiency are 1246 mg/g
similar to previous reports, we can see in Figure 2d that
and 62%, far less than that of nZVI@Mg(OH)2. Table 1 lists
unsupported nZVI particles tend to aggregate together, due to
van der Waals’ force and magnetic properties of the
Table 1. Comparison of Pb(II) Removal Performance of
particles.18,35 Enlarged TEM images (Figure S3) reveal that
nZVI@Mg(OH)2 with Other nZVI Based Materials
the nZVI particles have a diameter of approximately 25 nm,
Reported Previouslya
which is consistent with the calculated value through XRD
analysis. Moreover, a HAADF-STEM image of nZVI@Mg- specific surface area
(OH)2 and its corresponding EDS elemental mapping images (m2/g) capacity (mg/g)
are also presented in Figure 2e. Based on the O−K and Mg−K materials nZVI S.M. comp. nZVI S.M. comp.
signals and their distribution area, we find that the lamellar nZVI@Mg(OH)2 11.6 41.3 40.2 1718.4 775.4 1986.6
structure in designated area is Mg(OH)2 and Fe−K and Fe-L composite (this
study)
signal confirmed the particles belong to Fe (or its oxide), which
nZVI-Zeolite 12.3 1.0 83.4 806.0
certified the component of nZVI@Mg(OH)2 together with composite20
morphologic analysis.33 nZVI-Graphene 455.9 585.5
Pb(II) Removal from Water. The optimization experi- composite26
ments suggest that Pb(II) removal can be processed with a Sineguelas waste 3.9 35.6 63.5 225.0
supported nZVI36
neutral pH and 50% Fe content in composite (Figure S5).19
Kaolin supported 3.7 26.1 440.5
The study also shows that the initial Pb(II) concentration can nZVI17
barely influence the removal performance. Despite the fact that a
S.M. = support material, comp. = composite.
the initial concentration varies from 250 to 1000 mg/L, more
than 90% of Pb(II) is removed within 20 min for all samples.
Especially, when the initial concentration is lower than 750 mg/ the performance of some reported nZVI-based composite for
L, no Pb(II) is detected after 120 min (Figure S6). Pb(II) decontamination. Regardless of the varied experiment
Figure 3 presents the performance of Pb(II) removal by conditions in each study, to date, the nZVI@Mg(OH)2
Mg(OH)2, nZVI, and nZVI@Mg(OH)2 under the same composite in this work possesses the highest removal capacity.
optimal condition. As shown in Figure 3a, the Pb(II) In addition, we find that unsupported nZVI also exhibits a great
concentration decreases from the initial 1000 mg/L to 623.2, capacity (1718.4 mg/g) but still lower than that of nZVI@
141.8, and 6.8 mg/L after 120 min. The removal capacity of Mg(OH)2 composite. Therefore, we believe that the synergistic
Pb(II) for each material is calculated to be 775.4, 1718.4, and effect and good dispersion of nZVI particles on Mg(OH)2 can
1986.6 mg/g. The curves in Figure 3a show that the Pb(II) improve the removal performance effectively.18,20,26
Mechanisms of Pb(II) Removal by nZVI@Mg(OH)2. To
concentration decreases slowly for Mg(OH)2, while it occurs a
better understand the Pb(II) sequestration by the nZVI@
quick decline with fluctuant rate for unsupported nZVI. As for
Mg(OH)2 composite, the variations in phase, morphology,
nZVI@Mg(OH)2, the Pb(II) concentration drops more quickly
surface valence state of composite, and related metal ion
and ends with much lower residual concentration.16,18
concentrations were analyzed utilizing SEM/TEM, XRD, ICP-
Compared to Mg(OH)2 and nZVI, the removal performance AES, and high resolution X-ray photoelectron spectrometer
of the composite has obviously improved. To verify that the (HR-XPS).
improvement resulted from the synergistic effect of Mg(OH)2 Figure 4a illustrates the XRD pattern of the final solid sample
and nZVI, a weighted sum of capacity (as defined in formula 1) after treating 1000 mg/L Pb(II) solution. As shown in Figure
was calculated. 4a, the main phases are indexed as lead (Pb) and lead oxide
Q weightedsum = Q nZVI wt%nZVI + Q Mg(OH) wt%Mg(OH)2 (PbO), and there is no conspicuous peak attributed to nZVI
2 (1) and Mg(OH)2. Quantitative analyses by Rietveld spectrum
7964 DOI: 10.1021/am509184e
ACS Appl. Mater. Interfaces 2015, 7, 7961−7969
ACS Applied Materials & Interfaces Research Article

Figure 4. (a) XRD pattern of collected solid material after contacting with Pb(II) solution (dose, 500 mg/L; stirring speed, 300 r/min; contact time,
120 min; initial concentration, 1000 mg/L; initial pH, 6.86; temperature, 25 °C; ■, Pb; ●, PbO); (b) trends of mass fractions (normalized) of
relative elements in solid phase; (c)trends of concentrations of relative metal ions in liquid phase during the experiment.

fitting reveal a hybrid phase of 47% Pb and 53% PbO, (Fe 2p3/2), which would disappear after exposure to Pb(II)
suggesting that about half of Pb(II) is removed by reduction solution due to extensive oxidization.34,37 As shown in Figure
under the given circumstance. To retell the details of the 6b, the photoelectron spectrum for the O 1s region can be
sequestration process, variation of Pb, Fe, and Mg element decomposed into three peaks at 529.9, 531.2, and 532.5 eV,
distribution in both solid phase and liquid phase are measured which represent the binding energies of different oxygen in O−,
during batch experiments. As shown in Figure 4b, Pb mass OH−, and chemically or physically adsorbed water, respec-
fraction increases to over 90%, while Fe and Mg drop to 8.66% tively.33,37 According to the fitting data recorded by XPS PEAK
and 1.2% for the final solid sample. Combined with XRD software (Figure S8), the ratio of O−, OH−, and H2O varied
analysis, the finally obtained solids are mainly comprised of from 63.70%, 29.55%, and 6.75% to 49.54%, 42.68%, and
metallic lead and lead oxide. On the contrary, Figure 4c shows 7.78%, respectively, after reaction. The decrease of the O− ratio
that Pb(II) concentration in liquid samples decreases rapidly, and the increase of the OH− ratio suggest that the hydroxide,
while Fe(II/III) and Mg(II) concentrations rise to a certain including lead hydroxide and ferric hydroxide, is generated. In
level, due to the dissolution equilibrium of composite. Figure 6c, photoelectron peaks at 136.7 and 138.2 eV
Figure 5 exhibits the morphology of nZVI@Mg(OH)2 before correspond to Pb0 and PbO.19,37 We can see from the figure
and after reacting with Pb(II) solution. As shown in Figure that the PbO phase is dominant in the final solid samples, while
5c,d, the nZVI@Mg(OH)2 composite still presents a spherical Pb0 only accounts for 7.16%, which is different from that of the
shape and intersection structure after reacting with low XRD analyses. We believe that the limited penetration depth of
concentration (250 mg/L) Pb(II) solution, and the lamellar the XPS method and the susceptibility to oxidization of metallic
metallic lead or its compound is formed on the surface. lead are responsible for those differences.16,18
However, when the Pb(II) concentration is increased to 750 As the pH value can be a reliable tracer of the reactions
mg/L, the intersection structure is destroyed and replaced by during Pb(II) removal, the variation of the pH value was
lead and its compound packing together closely, in spite of the
recorded throughout the experiment. As shown in Figure 7a
remaining spherical shape of nZVI@Mg(OH)2 composite. The
(and Figure S7), without exception, a decline of pH emerges in
TEM micrograph (Figure 5b) also gives a clear version of the
all nZVI involved processes but is absent when dosed with
status of nZVI particles after being treated with low
Mg(OH)2 alone. This phenomenon can be explained by the
concentration (250 mg/L) Pb(II) solution, among which
some are intact and covered with thick-bedded target pollutant, hydroxylation of Fe(II/III), which can be derived from the
while others are corroded to be smaller-sized fragmentized following paths: (a) dissolution of iron oxides/hydroxides on
ones. the nZVI shell, (b) reaction between zerovalent iron and
HR-XPS was employed to investigate the surface state of Fe, water,38 and (c) reduction of Pb(II) by zerovalent iron.16,33
O, and Pb. Figure 6 presents a detailed XPS survey of Fe 2p, O Since the experiments were conducted under N2 environment,
1s, and Pb 4f regions. In Figure 6a, photoelectron peaks of the we presume that the anaerobic environment was well
Fe 2p region at 711 and 725 eV are attributed to the binding controlled, and the major reactions are listed as eqs 2−6.
energies of 2p3/2 and 2p1/2 of oxidized iron [Fe(III)], and the
weak peak at 706.7 eV suggests the existence of zerovalent iron Mg(OH)2 (s) → Mg 2 + + 2OH− (2)

7965 DOI: 10.1021/am509184e


ACS Appl. Mater. Interfaces 2015, 7, 7961−7969
ACS Applied Materials & Interfaces Research Article

Figure 5. (a) SEM image of nZVI@Mg(OH)2; (b) TEM image; (c and d) SEM images of collected solid material after contacting with low
concentration of Pb(II) solution (250 mg/L); (e and f) SEM images of collected solid material after contacting with high concentration of Pb(II)
solution (750 mg/L).

Pb2 + + 2OH− → Pb(OH)2 (s) (3) larger extent of pH drop (Figure S7), indicating that the pH
decline is indeed a consequence of Pb(II) reduction.
Pb2 + + Fe 0(s) → Fe 2 + + Pb0 (s) (4) Third, The dissolution of Mg(OH)2 continued (eq 2), and
the reaction between iron and water (eq 6) proceeded slowly
Fe2 + + 2H 2O → Fe(OH)2 (s) + 2H+ (5) and induced a jump of pH value. Therefore, Pb(II) could be
precipitated during this stage (eq 3).
Fe0(s) + 2H 2O → Fe 2 + + H 2(g) + 2OH− (6) The mechanism analyses suggest that Mg(OH)2 plays an
Based on the analysis above, a mechanism of Pb(II) removal essential role in Pb(II) immobilization for its dispersive,
by nZVI@Mg(OH)2 was proposed. In general, Pb(II) was supportive, and nZVI particle aggregate-preventive function as
sequestrated through a synergistic process, including adsorp- well as its intense inherent affinity to Pb(II). More significantly,
tion, reduction, ion exchange reaction, and subsequent a mutual stimulation between reactions 2 and 4 can occur; thus,
precipitation.16,18,19,33,37 According to the variation of the the hydroxies provided by Mg(OH)2 can dramatically promote
solution pH value, the removal of Pb(II) experienced three the role of nZVI as a reducer. In return, nZVI accelerates the
stages as shown in Figure 7b. dissolution of Mg(OH)2. The multifunction of Mg(OH)2
First, Pb(II) was attracted to the surface of nZVI particles makes it a good assistant for nZVI.
and Mg(OH)2, and then ion exchange reaction was initiated Remove Pb(II) from Acid Pb−Zn Mine Tailing
due to the dissolution of Mg(OH)2 (eq 2). At the same time, Leachate. The results of Pb(II) removal from acid Pb−Zn
Pb(II) was partly reduced to Pb0 by nZVI particles (eq 4), and mine tailing leachate by the nZVI@Mg(OH)2 composite are
the hydroxylation of Fe(II) (eq 5) would be triggered when it listed in Table 2. To meet the discharge standard (1.0 mg/L),
was accumulated to a certain concentration. The competition of we varied the dosage from 1000 to 10 000 mg/L and found that
above reactions resulted a slight ascendant of pH value. the residual concentration of Pb(II) and Zn(II) dropped to
Second, the reduction of Pb(II) became dominant, and thus 0.22 and 14.2 mg/L, respectively. When the dosage was 7000
the pH saw a rapid decline caused by the hydroxylation of mg/L, it corresponded to a removal efficiency of 99.9% and
Fe(II). pH changes of different initial Pb(II) concentrations 90.0%. The results certified that Pb(II) could be sequestrated
reveal that a higher initial Pb(II) concentration led to an earlier, effectively by this composite even under complicated solution
7966 DOI: 10.1021/am509184e
ACS Appl. Mater. Interfaces 2015, 7, 7961−7969
ACS Applied Materials & Interfaces Research Article

Figure 6. XPS survey of (a) the Fe 2p region, (b) the O 1s region of solid samples before and after treating Pb(II) solution (1000 mg/L), and (c)
XPS survey of the Pb 4f region of the final solid sample.

Figure 7. (a) Changes of solution pH during Pb(II) removal by different materials (initial Pb(II) concentration, 1000 mg/L); (b) stages of pH
variation during Pb(II) removal by nZVI@Mg(OH)2 (initial Pb(II) concentration, 1000 mg/L).

Table 2. Removal Performance Towards Acid Pb−Zn Mine


Tailing Leachate
■ CONCLUSIONS
In summary, a new Fe0 containing composite (nZVI@
Pb(II) Zn(II) pH Mg(OH)2) was prepared, characterized, and applied to remove
initial concentration (mg/L) 698.1 201.5 4.8
Pb(II). The morphology and phase analyses demonstrated that
after neutralized (mg/L) 412.0 142.8 7.0
nZVI particles were successfully loaded on the surface of the
ultimate concentration (mg/L) 0.22 14.2 7.2
self-supported flower-like Mg(OH)2. The results of Pb(II)
uptake amount (mg/g) 58.4 18.2 -
removal present a superb capacity (1986 mg/g) and large
removal efficiency (%) 99.9 90.0 -
kinetic removal rate (94% uptake within 15 min). We also
explored the mechanisms of Pb(II) deprivation, which revealed
that Mg(OH)2 can facilitate the sequestration by alleviating
aggregation of nZVI particles, enhancing Pb(II) reduction and
environment. Besides, the outlet pH value was relatively acting as an absorbent itself. The authors of this article believe
that it is the multirole of Mg(OH)2 and the synergistic effect
moderate compared to that made by the traditional alkali between nZVI and Mg(OH)2 that lead to the exceptional
precipitation method (as high as pH 10). performance of the composite. We think the results of this
7967 DOI: 10.1021/am509184e
ACS Appl. Mater. Interfaces 2015, 7, 7961−7969
ACS Applied Materials & Interfaces Research Article

work could provide an alternative technique for treating Pb(II)- (12) Li, X. Q.; Elliott, D. W.; Zhang, W. X. Zero-Valent Iron
bearing wastewater. Nanoparticles for Abatement of Environmental Pollutants: Materials


and Engineering Aspects. Crit. Rev. Solid State Mater. Sci. 2006, 31 (4),
ASSOCIATED CONTENT 111−122.
(13) Fu, F.; Dionysiou, D. D.; Liu, H. The Use of Zero-Valent Iron
*
S Supporting Information
for Groundwater Remediation and Wastewater Treatment: A Review.
Optimization of operation conditions, Pb(II) removal test in J. Hazard. Mater. 2014, 267, 194−205.
acid Pb−Zn mine tailing leachate, effect of initial concentration, (14) Fu, F.; Han, W.; Huang, C.; Tang, B.; Hu, M. Removal of
extra SEM and TEM images, and XPS fitting photoelectron Cr(VI) from Wastewater by Supported Nanoscale Zero-Valent Iron
spectra. This material is available free of charge via the Internet on Granular Activated Carbon. Desalin. Water Treat. 2013, 51 (13−
at http://pubs.acs.org. 15), 2680−2686.


(15) Kanel, S. R.; Manning, B.; Charlet, L.; Chol, H. Removal of
AUTHOR INFORMATION Arsenic(III) from Groundwater by Nanoscale Zero-Valent Iron.
Environ. Sci. Technol. 2005, 39, 1291−1298.
Corresponding Author (16) Ponder, S. M.; Darab, J. G.; Mallouk, T. E. Remediation of
*Tel.: +86 591 83792630. Fax: +86 591 83705474. E-mail: Cr(VI) and Pb(II) Aqueous Solutions Using Supported Nanoscale
zlin@fjirsm.ac.cn; zlin@scut.edu.cn. Zero-Valent Iron. Environ. Sci. Technol. 2000, 34 (12), 2564−2569.
Notes (17) Zhang, X.; Lin, S.; Lu, X. Q.; Chen, Z. l. Removal of Pb(II) from
The authors declare no competing financial interest. Water Using Synthesized Kaolin Supported Nanoscale Zero-Valent


Iron. Chem. Eng. J. 2010, 163 (3), 243−248.
(18) Zhang, X.; Lin, S.; Chen, Z.; Megharaj, M.; Naidu, R. Kaolinite-
ACKNOWLEDGMENTS Supported Nanoscale Zero-Valent Iron for Removal of Pb2+ from
Financial support was provided by the National Basic Research Aqueous Solution: Reactivity, Characterization and Mechanism. Water
Program of China (2013CB934302 and 2014CB932101), the Res. 2011, 45 (11), 3481−3488.
Outstanding Youth Fund (21125730), the “Strategic Priority (19) Zhang, Y.; Su, Y.; Zhou, X.; Dai, C.; Keller, A. A. A New Insight
Research Program” of the Chinese Academy of Sciences on the Core-Shell Structure of Zerovalent Iron Nanoparticles and Its
(XDA09030203), the National Science Foundation Grant Application for Pb(II) Sequestration. J. Hazard. Mater. 2013, 263 (Pt
(21477128), and the Technology Key Project of Fujian 2), 685−693.
Province (2013H0058). (20) Kim, S. A.; Kamala-Kannan, S.; Lee, K. J.; Park, Y. J.; Shea, P. J.;


Lee, W. H.; Kim, H. M.; Oh, B. T. Removal of Pb(II) from Aqueous
Solution by a Zeolite−Nanoscale Zero-Valent Iron Composite. Chem.
REFERENCES Eng. J. 2013, 217, 54−60.
(1) Nemsadze, K.; Sanikidze, T.; Ratiani, L.; Gabunia, L.; (21) Mueller, N. C.; Braun, J.; Bruns, J.; Cernik, M.; Rissing, P.;
Sharashenidze, T. Mechanisms of Lead-Induced Poisoning. Georgian Rickerby, D.; Nowack, B. Application of Nanoscale Zero Valent Iron
Med. News 2009, 1 (172−173), 92−96. (NZVI) for Groundwater Remediation in Europe. Environ. Sci. Pollut.
(2) Hou, S.; Yuan, L.; Jin, P.; Ding, B.; Qin, N.; Li, L.; Liu, X.; Wu, Res. Int. 2012, 19 (2), 550−558.
Z.; Zhao, G.; Deng, Y. A Clinical Study of the Effects of Lead (22) Phenrat, T.; Saleh, N.; Sirk, K.; Tilton, R. D.; Lowry, G. V.
Poisoning on the Intelligence and Neurobehavioral Abilities of Aggregation and Sedimentation of Aqueous Nanoscale Zerovalent
Children. Theor. Biol. Med. Modell. 2013, 10, 13. Iron Dispersions. Environ. Sci. Technol. 2007, 41 (1), 284−290.
(3) Yan, C. H.; Xu, J.; Shen, X. M. Childhood Lead Poisoning in (23) Yu, K.; Gu, C.; Boyd, S. A.; Liu, C.; Sun, C.; Teppen, B. J.; Li, H.
China: Challenges and Opportunities. Environ. Health Perspect. 2013, Rapid and Extensive Debromination of Decabromodiphenyl Ether by
121 (10), A294−A295. Smectite Clay-Templated Subnanoscale Zero-Valent Iron. Environ. Sci.
(4) Zhang, D.; Wang, M.; Tan, Y. L. Preparation of Porous Nano- Technol. 2012, 46 (16), 8969−8975.
barium-strontium Titanate by Sorghum Straw Template Method and (24) Zhang, Y.; Li, Y.; Li, J.; Hu, L.; Zheng, X. Enhanced Removal of
Its Adsorption Capability for Heavy Metal Ions. Chin. J. Chem. 2010, Nitrate by a Novel Composite: Nanoscale Zero Valent Iron Supported
68 (16), 1641−1648.
on Pillared Clay. Chem. Eng. J. 2011, 171 (2), 526−531.
(5) Fu, F.; Wang, Q. Removal of Heavy Metal Ions from
(25) Shi, L. N.; Zhang, X.; Chen, Z. L. Removal of Chromium (VI)
Wastewaters: A Review. J. Environ. Manage. 2011, 92 (3), 407−418.
from Wastewater Using Bentonite-Supported Nanoscale Zero-Valent
(6) Hua, M.; Zhang, S.; Pan, B.; Zhang, W.; Lv, L.; Zhang, Q. Heavy
Iron. Water Res. 2011, 45 (2), 886−892.
Metal Removal from Water/Wastewater by Nanosized Metal Oxides:
(26) Jabeen, H.; Kemp, K. C.; Chandra, V. Synthesis of Nano
A Review. J. Hazard. Mater. 2012, 211−212, 317−331.
(7) Crane, R. A.; Scott, T. B. Future Prospects for an Emerging Zerovalent Iron Nanoparticles-Graphene Composite for the Treat-
Water Treatment Technology. J. Hazard. Mater. 2012, 211−212, ment of Lead Contaminated Water. J. Environ. Manage. 2013, 130,
112−125. 429−435.
(8) Yan, W.; Lien, H. L.; Koel, B. E.; Zhang, W. X. Iron Nanoparticles (27) Liu, W.; Huang, F.; Wang, Y.; Zou, T.; Zheng, J.; Lin, Z.
for Environmental Clean-Up: Recent Developments and Future Recycling Mg(OH)2 Nanoadsorbent During Treating the Low
Outlook. Environ. Sci.: Processes Impacts 2013, 15 (1), 63−77. Concentration of CrVI. Environ. Sci. Technol. 2011, 45 (5), 1955−1961.
(9) Zheng, T. H.; Zhan, J. J.; He, J. B.; Day, C.; Lu, Y. F.; Mcpherson, (28) Wang, Y. J.; Chen, J.; Lu, L. L.; Lin, Z. Reversible Switch
G. L.; Piringer, G.; John, V. T. Reactivity Characteristics of Nanoscale between Bulk MgCO3·3H2O and Mg(OH)2 Micro/Nanorods Induces
Zerovalent Iron−Silica Composites for Trichloroethylene Remedia- Continuous Selective Preconcentration of Anionic Dyes. ACS Appl.
tion. Environ. Sci. Technol. 2008, 42 (12), 4494−4499. Mater. Interfaces 2013, 5 (16), 7698−7703.
(10) Ling, X.; Li, J.; Zhu, W.; Zhu, Y.; Sun, X.; Shen, J.; Han, W.; (29) Li, C. R.; Zhuang, Z. Y.; Huang, F.; Wu, Z. C.; Hong, Y. P.; Lin,
Wang, L. Synthesis of Nanoscale Zero-Valent Iron/Ordered Z. Recycling Rare Earth Elements from Industrial Wastewater with
Mesoporous Carbon for Adsorption and Synergistic Reduction of Flowerlike Nano-Mg(OH)2. ACS Appl. Mater. Interfaces 2013, 5 (19),
Nitrobenzene. Chemosphere 2012, 87 (6), 655−660. 9719−9725.
(11) Chen, Z.; Wang, T.; Jin, X.; Chen, Z.; Megharaj, M.; Naidu, R. (30) Estrada, M.; Costa, V. V.; Beloshapkin, S.; Fuentes, S.; Stoyanov,
Multifunctional Kaolinite-Supported Nanoscale Zero-Valent Iron Used E.; Gusevskaya, E. V.; Simakov, A. Aerobic Oxidation of Benzyl
for the Adsorption and Degradation of Crystal Violet in Aqueous Alcohol in Methanol Solutions over Au Nanoparticles: Mg(OH)2 Vs
Solution. J. Colloid Interface Sci. 2013, 398, 59−66. MgO as the Support. Appl. Catal., A 2014, 473, 96−103.

7968 DOI: 10.1021/am509184e


ACS Appl. Mater. Interfaces 2015, 7, 7961−7969
ACS Applied Materials & Interfaces Research Article

(31) Wang, Q. L.; Sndyder, S.; Kim, J.; Choi, A. Aqueous Ethanol
Modified Nanoscale Zerovalent Iron in Bromate Reduction, Synthesis,
Characterization, and Reactivity. Environ. Sci. Technol. 2009, 43 (9),
3292−3299.
(32) Sun, Y. P.; Li, X. Q.; Cao, J.; Zhang, W. X.; Wang, H. P.
Characterization of Zero-Valent Iron Nanoparticles. Adv. Colloid
Interface Sci. 2006, 120 (1−3), 47−56.
(33) Xi, Y.; Mallavarapu, M.; Naidu, R. Reduction and Adsorption of
Pb2+ in Aqueous Solution by Nano-Zero-Valent Irona SEM, TEM
and XPS Study. Mater. Res. Bull. 2010, 45 (10), 1361−1367.
(34) Li, X. Q.; Zhang, W. X. Iron Nanoparticles, the Core-Shell
Structure and Unique Properties for Ni(II) Sequestration. Langmuir
2006, 22, 4638−4642.
(35) Li, L.; Fan, M. H.; Brown, R. C.; Leeuwen, J. H. V.; Wang, J.;
Wang, W.; Song, Y.; Zhang, P. Synthesis, Properties, and Environ-
mental Applications of Nanoscale Iron-Based Materials: A Review.
Crit. Rev. Env. Sci. Technol. 2010, 36 (5), 405−431.
(36) Arshadi, M.; Soleymanzadeh, M.; Salvacion, J. W.; SalimiVahid,
F. Nanoscale Zero-Valent Iron (NZVI) Supported on Sineguelas
Waste for Pb(II) Removal from Aqueous Solution: Kinetics,
Thermodynamic and Mechanism. J. Colloid Interface Sci. 2014, 426,
241−251.
(37) Li, X. Q.; Zhang, W. X. Sequestration of Metal Cations with
Zerovalent Iron Nanoparticles−A Study with High Resolution X-ray
Photoelectron Spectroscopy (HR-XPS). J. Phys. Chem. C 2007, 11,
6939−6946.
(38) Filip, J.; Karlický, F.; Marušaḱ , Z.; Lazar, P.; Č erník, M.;
Otyepka, M.; Zbořil, R. Anaerobic Reaction of Nanoscale Zerovalent
Iron with Water: Mechanism and Kinetics. J. Phys. Chem. C 2014, 118
(25), 13817−13825.

7969 DOI: 10.1021/am509184e


ACS Appl. Mater. Interfaces 2015, 7, 7961−7969

You might also like